首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 203 毫秒
1.
The conformation of the nitroxyl spin-label ethyl 3-(2,2,5,5-tetramethylpyrrolinyl-1-oxyl)-propen-2-oate has been determined by electron nuclear double resonance (ENDOR) spectroscopy and computer-based molecular modeling. From ENDOR spectra of the compound in frozen solution, we have assigned resonance absorption features for each class of protons, and we have identified their principal hyperfine coupling (hfc) components from analysis of the dependence of ENDOR spectra on the static laboratory magnetic field. The dipolar hfc components yielded estimates of the electron-proton separations for each class of protons of the ethyl propenoyl moiety. Torsion angle search calculations were carried out to determine the conformational space compatible with hard-sphere nonbonded constraints and with the ENDOR-determined distance constraints. Molecular graphics analysis revealed that the propenoyl side chain of the spin-label exhibits an extended trans conformation and that the ethyl moiety of the ester group deviates significantly from coplanarity with the carboxylate--COO--atoms. The conformation of this molecule is compared with that of an analogous compound O-[3-(2,2,5,5-tetramethylpyrrolinyl-1-oxyl)-propen-2-oyl]-L- beta- phenyllactate, which has been employed as a spectroscopic substrate probe of carboxypeptidase A (L. C. Kuo, J. M. Fukuyama, and M. W. Makinen (1983) Journal of Molecular Biology 163, 63-105). The rotamer conformation of the free spin-label ester in solution, as determined in this study, and that of the enzyme-bound spin-labeled phenyllactate are compared. Differences in rotamer structure are discussed in terms of stereoelectronic principles that govern the pathway of substrate hydrolysis catalyzed by carboxypeptidase A.  相似文献   

2.
The catalytically competent active-site structure of a true acylenzyme reaction intermediate of TEM-1 beta-lactamase formed with the kinetically specific spin-labeled substrate 6-N-(2,2,5,5-tetramethyl-1-oxypyrrolinyl-3-carboxyl)-penicillanic acid isolated under cryoenzymologic conditions has been determined by angle-selected electron nuclear double resonance (ENDOR) spectroscopy. Cryoenzymologic experiments with use of the chromophoric substrate 6-N-[3-(2-furanyl)-propen-2-oyl]-penicillanic acid showed that the acylenzyme reaction intermediate could be stabilized in the -35 to -75 degrees C range with a half-life suitably long to allow freeze-quenching of the reaction species for ENDOR studies while a noncovalent Michaelis complex could be optically identified at temperatures only below -70 degrees C. The wild-type, Glu166Asn, Glu240Cys, and Met272Cys mutant forms of the mature enzyme were overexpressed in perdeuterated minimal medium to allow detection and assignment of proton resonances specific for the substrate and chemically modified amino acid residues in the active site. From analysis of the dependence of the ENDOR spectra on the setting of the static laboratory magnetic field H0, the dipolar contributions to the principal hyperfine coupling components were estimated to calculate the separations between the unpaired electron of the nitroxyl group and isotopically identified nuclei. These electron-nucleus distances were applied as constraints to assign the conformation of the substrate in the active site and of amino acid side chains by molecular modeling. Of special interest was that the ENDOR spectra revealed a water molecule sequestered in the active site of the acylenzyme of the wild-type protein that was not detected in the deacylation impaired Glu166Asn mutant. On the basis of the X-ray structure of the enzyme, the ENDOR distance constraints placed this water molecule within hydrogen-bonding distance to the carboxylate side chain of glutamate-166 as if it were poised for nucleophilic attack of the scissile ester bond. The ENDOR results provide experimental evidence of glutamate-166 in its functional role as the general base catalyst in the wild-type enzyme for hydrolytic breakdown of the acylenzyme reaction intermediate of TEM-1 beta-lactamase.  相似文献   

3.
The enzyme 6-pyruvoyl tetrahydropterin synthase (PTPS) catalyses the second step in the de novo biosynthesis of tetrahydrobiopterin, the conversion of dihydroneopterin triphosphate to 6-pyruvoyl tetrahydropterin. The Zn and Mg-dependent reaction includes a triphosphate elimination, a stereospecific reduction of the N5-C6 double bond and the oxidation of both side-chain hydroxyl groups. The crystal structure of the inactive mutant Cys42Ala of PTPS in complex with its natural substrate dihydroneopterinetriphosphate was determined at 1.9 A resolution. Additionally, the uncomplexed enzyme was refined to 2.0 A resolution. The active site of PTPS consists of the pterin-anchoring Glu A107 neighboured by two catalytic motifs: a Zn(II) binding site and an intersubunit catalytic triad formed by Cys A42, Asp B88 and His B89. In the free enzyme the Zn(II) is in tetravalent co-ordination with three histidine ligands and a water molecule. In the complex the water is replaced by the two substrate side-chain hydroxyl groups yielding a penta-co-ordinated Zn(II) ion. The Zn(II) ion plays a crucial role in catalysis. It activates the protons of the substrate, stabilizes the intermediates and disfavours the breaking of the C1'C2' bond in the pyruvoyl side-chain. Cys A42 is activated by His B89 and Asp B88 for proton abstraction from the two different substrate side-chain atoms C1', and C2'. Replacing Ala A42 in the mutant structure by the wild-type Cys by modelling shows that the C1' and C2' substrate side-chain protons are at equal distances to Cys A42 Sgamma. The basicity of Cys A42 may be increased by a catalytic triad His B89 and Asp B88. The active site of PTPS seems to be optimised to carry out proton abstractions from two different side-chain C1' and C2' atoms, with no obvious preference for one of them. Kinetic studies with dihydroneopterin monophosphate reveal that the triphosphate moiety of the substrate is necessary for enzyme specifity.  相似文献   

4.
The conformations of enzyme-bound pentapeptide (Arg-Arg-Ala-Ser-Leu) and heptapeptide (Leu-Arg-Arg-Ala-Ser-Leu-Gly) substrates of protein kinase have been studied by NMR in quaternary complexes of the type (Formula: see text). Paramagnetic effects of Mn2+ bound at the inhibitory site of the catalytic subunit on the longitudinal relaxation rates of backbone Ca protons, as well as on side-chain protons of the bound pentapeptide and heptapeptide substrates, have been used to determine Mn2+ to proton distances which range from 8.2 to 12.4 A. A combination of the paramagnetic probe-T1 method with the Redfield 2-1-4-1-2 pulse sequence for suppression of the water signal has been used to measure distances from Mn2+ to all of the backbone amide (NH) protons of the bound pentapeptide and heptapeptide substrates, which range from 6.8 to 11.1 A. Paramagnetic effects on the transverse relaxation rates yield rate constants for peptide exchange, indicating that the complexes studied by NMR dissociate rapidly enough to participate in catalysis. Model-building studies based on the Mn2+-proton distances, as well as on previously determined distances from Cr3+-AMPPCP to side-chain protons [Granot, J., Mildvan, A.S., Bramson, H. N., & Kaiser, E. T. (1981) Biochemistry 20, 602], rule out alpha-helical, beta-sheet, beta-bulge, and all possible beta-turn conformations within the bound pentapeptide and heptapeptide substrates. The distances are fit only by extended coil conformations for the bound peptide substrates with a minor difference between the pentapeptides and heptapeptides in the phi torsional angle at Arg3C alpha and in psi at Arg2C alpha. An extended coil conformation, which minimizes the number of interactions within the substrate, would facilitate enzyme-substrate interaction and could thereby contribute to the specificity of protein kinase.  相似文献   

5.
The hydrolysis of beta-lactam antibiotics by class A beta-lactamases is a common cause of bacterial resistance to these agents. The beta-lactamase inhibitory protein (BLIP) is able to bind and inhibit several class A beta-lactamases, including TEM-1 beta-lactamase and SME-1 beta-lactamase. Although the TEM-1 and SME-1 enzymes share 33% amino acid sequence identity and a similar fold, they differ substantially in surface electrostatic properties and the conformation of a loop-helix region that BLIP binds. Alanine-scanning mutagenesis was performed to identify the residues on BLIP that contribute to its binding affinity for each of these enzymes. The results indicate that the sequence requirements for binding are similar for both enzymes with most of the binding free energy provided by two patches of aromatic residues on the surface of BLIP. Polar residues such as several serines in the interface do not make significant contributions to affinity for either enzyme. In addition, the specificity of binding is significantly altered by mutation of two charged residues, Glu73 and Lys74, that are buried in the structure of the TEM-1.BLIP complex as well as by residues located on two loops that insert into the active site pocket. Based on the results, a E73A/Y50A double mutant was constructed that exhibited a 220,000-fold change in binding specificity for the TEM-1 versus SME-1 enzymes.  相似文献   

6.
F Ni  D R Ripoll  E O Purisima 《Biochemistry》1992,31(9):2545-2554
The COOH-terminal region of hirudin represents an independent functional domain that binds to an anion-binding exosite of thrombin and inhibits the interaction of thrombin with fibrinogen and regulatory proteins in blood coagulation. The thrombin-bound structure of the peptide fragment, hirudin 55-65, has been determined by use of transferred NOE spectroscopy [Ni, F., Konishi, Y., & Scheraga, H. A. (1990) Biochemistry 29, 4479-4489]. The stability of the thrombin-bound conformation has been characterized further by a combined NMR and theoretical analysis of the conformational ensemble accessible by the hirudin peptide. Medium- and long-range NOE's were found for the free hirudin peptide in aqueous solution and in a mixture of dimethyl sulfoxide and water at both ambient (25 degrees C) and low (0 degrees C) temperatures, suggesting that ordered conformations are highly populated in solution. The global folding of these conformations is similar to that in the thrombin-bound state, as indicated by NOE's involving the side-chain protons of residues Phe(56), Ile(59), Pro(60), Tyr(63), and Leu(64). Residues Glu(61), Glu(62), Tyr(63), and Leu(64) all contain approximately 50% of helical conformations calculated from the ratio of the sequential dNN and d alpha N NOE's. Among the helical ensemble, active 3(10)-helical conformations were found by an analysis of the medium-range [(i,i+2) and (i,i+3)] NOE's involving the last six residues of the peptide. An analysis of the side-chain rotamers revealed that, upon binding to thrombin, there may be a rotation around the alpha CH-beta CH bond of Ile(59) such that Ile(59) adopts a gauche- (chi 1 = +60) conformation in contrast to the highly populated trans (chi 1 = -60) found for Ile(59) in the free peptide. However, the thrombin-bound conformation of the hirudin peptide is still an intrinsically stable conformer, and the preferred conformational ensemble of the peptide contains a large population of the active conformation. The apparent preference for a gauche- (chi 1 = +60) side-chain conformation of Ile(59) in the bound state may be explained by the existence of a positively charged arginine residue among the hydrophobic residues in the thrombin exosite.  相似文献   

7.
M Zhao  K C Zen  W L Hubbell  H R Kaback 《Biochemistry》1999,38(23):7407-7412
Evidence has been presented [Venkatesan, P., and Kaback, H. R. (1998) Proc. Natl. Acad. Sci. U.S.A. 95, 9802-9807] that Glu126 (helix IV) and Arg144 (helix V) which are critical for substrate binding in the lactose permease of Escherichia coli are charge paired and therefore in close proximity. To test this conclusion more directly, three different site-directed spectroscopic techniques were applied to permease mutants in which Glu126 and/or Arg144 were replaced with either His or Cys residues. (1) Glu126-->His/Arg144-->His permease containing a biotin acceptor domain was purified by monomeric avidin affinity chromatography, and Mn(II) binding was assessed by electron paramagnetic resonance spectroscopy. The mutant protein binds Mn(II) with a KD of about 40 microM at pH 7.5, while no binding is observed at pH 5.5. In addition, no binding is detected with Glu126-->His or Arg144-->His permease. (2) Permease with Glu126-->Cys/Arg144-->Cys and a biotin acceptor domain was purified, labeled with a thiol-specific nitroxide spin-label, and shown to exhibit spin-spin interactions in the frozen state after reconstitution into proteoliposomes. (3) Glu126-->Cys/Arg144-->Cys permease with a biotin acceptor domain was purified and labeled with a thiol-specific pyrene derivative, and fluorescence spectra were obtained after reconstitution into lipid bilayers. An excimer band is observed with the reconstituted E126C/R144C mutant, but not with either single-Cys mutant or when the single-Cys mutants are mixed prior to reconstitution. The results provide strong support for the conclusion that Glu126 (helix IV) and Arg144 (helix V) are in close physical proximity.  相似文献   

8.
Vicinal coupling constants between various nuclei provide backbone and side-chain conformational information for a series of asparagine- and tyrosine-containing peptides in DMSO and in H2O. By enriching Tyr of Ac-Asn-Pro-Tyr-NHMe with 15N, it has been possible to distinguish between the resonances of the two side-chain beta protons of Tyr. Analysis of the coupling constants in terms of the distributions of side-chain conformations in these peptides indicates that the addition of Asn to the Pro-Tyr sequence leads to a less random conformational distribution. When compared to the side-chain rotamer distribution of Ac-Asn-NHMe and Ac-Tyr-NHMe, particular Asn and Tyr side-chain conformations of Ac-Asn-Pro-Tyr-NHMe are stabilized in dimethylsulfoxide solution. The interaction(s) which stabilize a unique Tyr side-chain conformation of Ac-Asn-Pro-Tyr-NHMe in dimethylsulfoxide are not present in Ac-Ala-Pro-Tyr-NHMe and are unaffected by the addition of Val-Pro to the C-terminus of Asn-Pro-Tyr. In water, a preferential stabilization of one Asn side-chain conformation of Ac-Asn-Pro-Tyr-NHMe is also observed, while the Tyr side-chain rotamer distribution is similar to that of Ac-Tyr-NHMe. An interaction between the Asn side chain and the Pro-Tyr-NHMe backbone was previously shown to stabilize a beta-bend conformation at Pro-Tyr in water. Data are also presented for Ac-Tyr-Pro-Asn-NHMe, for which local interactions do not stabilize particular backbone conformations in dimethylsulfoxide or in water. The conformations of the peptides studied here are relatively insensitive to temperatures between 27 degrees and 62 degrees, both in dimethylsulfoxide and in water. The sequences Asn-Pro-Tyr and Tyr-Pro-Asn occur in ribonuclease A, and these tripeptides serve as models for the interactions involved in the folding of this protein.  相似文献   

9.
An expression system has been developed that allows high levels of production of TEM-1 beta-lactamase with ease of biosynthetic incorporation of nuclear isotopes. The gene for mature TEM-1 beta-lactamase fused to the leader sequence of the ompA protein was subcloned into the pET-24a(+) vector by introduction of an NdeI restriction site at the first codon of the fused genes and transformed into Escherichia coli BL21 (DE3) cells. With protein induction at 25 degrees C supported by LB medium supplemented with osmolytes (300 mM sucrose and 2.5 mM betaine), the extracellular, mature form of wild-type TEM-1 beta-lactamase was recovered at a level of 140 mg/L. The production level of E166N, E240C, E104C, and M272C mutants depended on the mutation but was invariably higher than reported by others for expression systems of the wild-type enzyme. Comparison of different carbon sources on the efficiency of biosynthetic incorporation of covalent deuterium showed maximal (90%) incorporation with minimal medium containing 99% (2)H(2)O and sodium d(3)-acetate (99 atom% (2)H). The yield of deuterium-enriched wild-type enzyme was 80 mg/L with yields for mutants proportionally reduced. The high level of protein deuteration achieved with this system allowed detection of the hyperfine coupling between the paramagnetic nitroxyl group of a spin-labeled penicillin substrate and hydrogens on the penicillin moiety in a cryokinetically isolated acylenzyme reaction intermediate because of the decrease in overlapping resonances of active site residues. The overexpression system is readily adaptable for other target proteins and facilitates studies requiring large quantities of protein in isotopically enriched forms.  相似文献   

10.
Beta-lactamase confers resistance to penicillin-like antibiotics by hydrolyzing their beta-lactam bond. To combat these enzymes, inhibitors covalently cross-linking the hydrolytic Ser70 to Ser130 were introduced. In turn, mutant beta-lactamases have emerged with decreased susceptibility to these mechanism-based inhibitors. Substituting Ser130 with glycine in the inhibitor-resistant TEM (IRT) mutant TEM-76 (S130G) prevents the irreversible cross-linking step. Since the completely conserved Ser130 is thought to transfer a proton important for catalysis, its substitution might be hypothesized to result in a nonfunctional enzyme; this is clearly not the case. To investigate how TEM-76 remains active, its structure was determined by X-ray crystallography to 1.40 A resolution. A new water molecule (Wat1023) is observed in the active site, with two configurations located 1.1 and 1.3 A from the missing Ser130 Ogamma; this water molecule likely replaces the Ser130 side-chain hydroxyl in substrate hydrolysis. Intriguingly, this same water molecule is seen in the IRT TEM-32 (M69I/M182T), where Ser130 has moved significantly. TEM-76 shares other structural similarities with various IRTs; like TEM-30 (R244S) and TEM-84 (N276D), the water molecule activating clavulanate for cross-linking (Wat1614) is disordered (in TEM-30 it is actually absent). As expected, TEM-76 has decreased kinetic activity, likely due to the replacement of the Ser130 side-chain hydroxyl with a water molecule. In contrast to the recently determined structure of the S130G mutant in the related SHV-1 beta-lactamase, in TEM-76 the key hydrolytic water (Wat1561) is still present. The conservation of similar accommodations among IRT mutants suggests that resistance arises from common mechanisms, despite the disparate locations of the various substitutions.  相似文献   

11.
The positively charged quaternary ammonium group of agonists of the nicotinic acetylcholine (ACh) receptor binds to a negative subsite at most about 1 nm from a readily reducible disulfide. This disulfide is formed by alpha Cys192 and Cys193 (Kao and Karlin, 1986). In order to identify Asp or Glu residues that may contribute to the negative subsite, we synthesized S-(2-[3H]glycylamidoethyl)dithio-2-pyridine. Purified ACh receptor from Torpedo californica was mildly reduced and reacted with S-(2-[3H]glycylamidoethyl)dithio-2-pyridine. The predominant product was a mixed disulfide between the 3H-N-glycylcysteamine moiety and alpha Cys192 or Cys193. In the extended conformation of [3H] N-glycylcysteamine, the distance from the glycyl amino group to the cysteamine thio group is 0.9 nm. Thus, the amino group of disulfide-linked [3H]N-glycylcysteamine could react with carboxyls within 0.9 nm of Cys192/Cys193. To promote amide bond formation between the tethered amino group and receptor carboxyls, we added 1-ethyl-3-(3'-dimethylaminopropyl)-carbodiimide. The predominant sites of amide coupling were on the delta subunit, in CNBr fragment 4 (delta 164-257). This reaction was inhibited by ACh. Only the first 61 residues of delta CNBr 4 are predicted to be extracellular, and there are 11 Asp or Gly residues in this region. One or more of these residues is likely to contribute to the binding of ACh.  相似文献   

12.
A combined 1H-NMR and molecular mechanics study of [Cpp1, Sar7]AVP was performed in order to select the most probable conformations in DMSO solutions. The NMR constraints obtained were employed in the selection of starting conformations of the cyclic moiety of the analog. In particular, the diminished accessibility of the Asn5 NH proton to solvent and the close contact between Cpp1 and Cys6 C alpha H protons suggests a beta-turn conformation at the Phe3-Gln4 residues. Energy minimization was carried out both in the ECEPP/2 (rigid-valence geometry) and in the AMBER (flexible-valence geometry) force fields. Comparison of the experimental and calculated values of NMR characteristics has revealed that conformations containing type I, II, and III beta-turns at the Phe3-Gln4 residues are in reasonable agreement with the experimental data, with a dynamic equilibrium between the beta I (beta III) and beta II type structures of the cyclic part being the most probable. All of these conformations prefer the negative chirality of the disulfide bridge (theta 3 approximately -90 degrees). Five representative conformations were chosen for the acyclic tail: one with a beta I, one with a beta II'-turn at the Sar7-Arg8 residues, two extended-type conformations, and a conformation with a gamma-turn at Sar7. Because only high-energy extended conformations were in agreement with NMR data, it was concluded that the acyclic tail has considerable conformational flexibility in solution. The conformations obtained are discussed in terms of the structure-function relationship of the neurohypophyseal hormone analogs.  相似文献   

13.
Simulation studies have been performed to evaluate the utility of site-directed spin labeling for determining the structures of protein-ligand complexes, given a known protein structure. Two protein-ligand complexes were used as model systems for these studies: a 1.9-A-resolution x-ray structure of a dihydrofolate reductase mutant complexed with methotrexate, and a 1.5-A-resolution x-ray structure of the V-Src tyrosine kinase SH2 domain complexed with a five-residue phosphopeptide. Nitroxide spin labels were modeled at five dihydrofolate reductase residue positions and at four SH2 domain residue positions. For both systems, after energy minimization, conformational ensembles of the spin-labeled residues were generated by simulated annealing while holding the remainder of the protein-ligand complex fixed. Effective distances, simulating those that could be obtained from (1)H-NMR relaxation measurements, were calculated between ligand protons and the spin labels. These were converted to restraints with several different levels of precision. Restrained simulated annealing calculations were then performed with the aim of reproducing target ligand-binding modes. The effects of incorporating a few supplementary short-range (< or =5.0 A) distance restraints were also examined. For the dihydrofolate reductase-methotrexate complex, the ligand-binding mode was reproduced reasonably well using relatively tight spin-label restraints, but methotrexate was poorly localized using loose spin-label restraints. Short-range and spin-label restraints proved to be complementary. For the SH2 domain-phosphopeptide complex without the short-range restraints, the peptide did not localize to the correct depth in the binding groove; nevertheless, the orientation and internal conformation of the peptide was reproduced moderately well. Use of the spin-label restraints in conjunction with the short-range restraints resulted in relatively well defined structural ensembles. These results indicate that restraints derived from site-directed spin labeling can contribute significantly to defining the orientations and conformations of bound ligands. Accurate ligand localization appears to require either a few supplementary short-range distance restraints, or relatively tight spin-label restraints, with at least one spin label positioned so that some of the restraints draw the ligand into the binding pocket in the latter case.  相似文献   

14.
X-ray crystal structures of enzyme-ligand complexes are widely believed to mimic states in the catalytic cycle, but this presumption has seldom been carefully scrutinized. In the case of Tritrichomonas foetus inosine 5'-monophosphate dehydrogenase (IMPDH), 10 structures of various enzyme-substrate-inhibitor complexes have been determined. The Cys319 loop is found in at least three different conformations, suggesting that its conformation changes as the catalytic cycle progresses from the dehydrogenase step to the hydrolase reaction. Alternatively, only one conformation of the Cys319 loop may be catalytically relevant while the others are off-pathway. Here we differentiate between these two hypotheses by analyzing the effects of Ala substitutions at three residues of the Cys319 loop, Arg322, Glu323, and Gln324. These mutations have minimal effects on the value of k(cat) (≤5-fold) that obscure large effects (>10-fold) on the microscopic rate constants for individual steps. These substitutions increase the equilibrium constant for the dehydrogenase step but decrease the equilibrium between open and closed conformations of a mobile flap. More dramatic effects are observed when Arg322 is substituted with Glu, which decreases the rates of hydride transfer and hydrolysis by factors of 2000 and 130, respectively. These experiments suggest that the Cys319 loop does indeed have different conformations during the dehydrogenase and hydrolase reactions as suggested by the crystal structures. Importantly, these experiments reveal that the structure of the Cys319 loop modulates the closure of the mobile flap. This conformational change converts the enzyme from a dehydrogenase into hydrolase, suggesting that the conformation of the Cys319 loop may gate the catalytic cycle.  相似文献   

15.
Glycosylphosphatidylinositol-specific phospholipase C (GPtdIns-PLC) is found in the protozoan parasite Trypanosoma brucei. A region of protein sequence similarity exists between the protozoan enzyme and eubacterial phosphatidylinositol-phospholipases C. The functional relevance of Cys80 and Gln81 of GPtdIns-PLC, both in this region, was tested with a panel of mutations at each position. Gln81Glu, Gln81Ala, Gln81Gly, Gln81Lys and Gln81Leu mutants were inactive. Cleavage of GPtdIns was detectable in Gln81Asn, although the specific activity decreased 500-fold, and kcat was reduced 50-fold. Thus an amide side-chain at residue 81 is essential for catalysis by GPtdIns-PLC. Sulfhydryl reagents inactivate GPtdIns-PLC, suggesting that a Cys could be close to the enzyme active site. Surprisingly, p-chloromercuriphenyl sulfonate (p-CMPS) is significantly more potent than N-ethylmaleimide, the less bulky compound. This knowledge prompted us to test whether replacement of Cys80 with an amino acid possessing a bulky side-chain would inactivate GPtdIns-PLC: Cys80Ala, Cys80Thr, Cys80Phe, Cys184Ala, and Cys269-270-273Ser were constructed for that purpose. Cys80Phe lacked enzyme activity, while Cys80Ala, Cys80Thr and Cys269-270-273Ser retained 33 to 100% of wild-type activity. Interestingly, the Cys80Ala and Cys80Thr mutants became resistant to p-CMPS, as predicted if the sulfhydryl reagent reacted with Cys80 in the wild-type enzyme to form a cysteinyl mercurylphenylsulfonate moiety, a bulky adduct that inactivated GPtdIns-PLC, similar to the Cys80Phe mutation. We conclude that a bulky side-chain (or adduct) at position 80 of GPtdIns-PLC abolishes enzyme activity. Together, these observations place Cys80 and Gln81 at, or close to, the active site of GPtdIns-PLC from T. brucei.  相似文献   

16.
The R-specific alcohol dehydrogenase (RADH) from Lactobacillus brevis is an NADP-dependent, homotetrameric member of the extended enzyme family of short-chain dehydrogenases/reductases (SDR) with a high biotechnological application potential. Its preferred in vitro substrates are prochiral ketones like acetophenone with almost invariably a small methyl group as one substituent and a bulky (often aromatic) moiety as the other. On the basis of an atomic-resolution structure of wild-type RADH in complex with NADP and acetophenone, we designed the mutant RADH-G37D, which should possess an improved cosubstrate specificity profile for biotechnological purposes, namely, a preference for NAD rather than NADP. Comparative kinetic measurements with wild-type and mutant RADH showed that this aim was achieved. To characterize the successful mutant structurally, we determined several, partly atomic-resolution, crystal structures of RADH-G37D both as an apo-enzyme and as ternary complex with NAD or NADH and phenylethanol. The increased affinity of RADH-G37D for NAD(H) depends on an interaction between the adenosine ribose moiety of NAD and the inserted aspartate side-chain. A structural comparison between RADH-G37D as apo-enzyme and as a part of a ternary complex revealed significant rearrangements of Ser141, Glu144, Tyr189 and Met205 in the vicinity of the active site. This plasticity contributes to generate a small hydrophobic pocket for the methyl group typical for RADH substrates, and a hydrophobic coat for the second, more variable and often aromatic, substituent. Around Ser141 we even found alternative conformations in the backbone. A structural adaptability in this region, which we describe here for the first time for an SDR enzyme, is probably functionally important, because it concerns Ser142, a member of the highly conserved catalytic tetrad typical for SDR enzymes. Moreover, it affects an extended proton relay system that has been identified recently as a critical element for the catalytic mechanism in SDR enzymes.  相似文献   

17.
2D 1H-NMR spectra of des-Gly9-[Arg8]vasopressin in dimethylsulfoxide have been taken and the 1H resonances have been assigned. The coupling constants and amide proton temperature coefficients (delta delta/delta T) have been measured and the NOE cross-peaks in the NOESY spectrum have been analyzed. The most essential information on the spatial structure of des-Gly9-[Arg8]vasopressin is extracted from the low delta delta/delta T value for Asn5 amide proton and from the NOE between the Cys1 and Cys6 alpha-protons. A diminished accessibility of the Asn5 NH proton for the solvent is ascribed to the presence of a beta-turn in the fragment 2-5. The distance between the Cys1 and Cys6 C alpha H protons seems to be less than 4 A. These constraints were taken into account in the conformational analysis of the title peptide. The derived set of the low-energy backbone conformations was analyzed against the background of the all available NMR data. The most probable conformation of the cyclic moiety in des-Gly9-[Arg8]vasopressin was found to be the type III beta-turn. The corner positions are occupied by the residues 3, 4, while the residues 1-2 and 5-6 are at the extended sites. Some NMR data indicate that this structure is in a dynamic equilibrium with other minor conformers.  相似文献   

18.
Hauser K  Barth A 《Biophysical journal》2007,93(9):3259-3270
Protonation of acidic residues in the sarcoplasmic reticulum Ca(2+)-ATPase (SERCA 1a) was studied by multiconformation continuum electrostatic calculations in the Ca(2+)-bound state Ca(2)E1, in the Ca(2+)-free state E2(TG) with bound thapsigargin, and in the E2P (ADP-insensitive phosphoenzyme) analog state with MgF(4)(2-) E2(TG+MgF(4)(2-)). Around physiological pH, all acidic Ca(2+) ligands (Glu(309), Glu(771), Asp(800), and Glu(908)) were unprotonated in Ca(2)E1; in E2(TG) and E2(TG+MgF(4)(2-)) Glu(771), Asp(800), and Glu(908) were protonated. Glu(771) and Glu(908) had calculated pK(a) values larger than 14 in E2(TG) and E2(TG+MgF(4)(2-)), whereas Asp(800) titrated with calculated pK(a) values near 7.5. Glu(309) had very different pK(a) values in the Ca(2+)-free states: 8.4 in E2(TG+MgF(4)(2-)) and 4.7 in E2(TG) because of a different local backbone conformation. This indicates that Glu(309) can switch between a high and a low pK(a) mode, depending on the local backbone conformation. Protonated Glu(309) occupied predominantly two main, very differently orientated side-chain conformations in E2(TG+MgF(4)(2-)): one oriented inward toward the other Ca(2+) ligands and one oriented outward toward a protein channel that seems to be in contact with the cytoplasm. Upon deprotonation, Glu(309) adopted completely the outwardly orientated side-chain conformation. The contact of Glu(309) with the cytoplasm in E2(TG+MgF(4)(2-)) makes this residue unlikely to bind lumenal protons. Instead it might serve as a proton shuttle between Ca(2+)-binding site I and the cytoplasm. Glu(771), Asp(800), and Glu(908) are proposed to take part in proton countertransport.  相似文献   

19.
20.
Single-chain Fv (scFv) antibody libraries were constructed from mice immunized with an ampicillin-bovine serum albumin conjugate. Several antibodies with specificity for intact ampicillin were selected by phage display and characterized. The antibody scFv fragment aL2 binds to intact ampicillin and shows no detectable cross-reactivity with hydrolyzed ampicillin. We determined the X-ray structures of two crystal forms of w.t. aL2, which differ mainly in the side-chain conformation of Trp H109 (according to a new consensus nomenclature Kabat residue number H95) in the extremely short (three residues) CDR H3 and the presence or absence of a well-resolved molecule of 2-methyl-pentane-2,4-diol in the bottom of the binding pocket. Attempts to co-crystallize aL2 with its antigen or to diffuse ampicillin into the wild-type aL2 crystals were unsuccessful, since crystal contacts obstruct the binding pocket. However, a mutant with two point mutations near the N terminus (Gln H6 replaced by Glu and Ala H10 (Kabat H9) replaced by Gly) crystallized in a form compatible with antigen-binding. Although the mutations affect the conformation of framework I, the conformations of the binding pocket of the uncomplexed wild-type aL2 and of the mutant complex were almost identical. The structure explains the specificity of the antibody for intact ampicillin and the degree of cross-reactivity of aL2 with a wide variety of ampicillin analogs. This antibody system will be very useful as a diagnostic reagent for antibiotics use and abuse, as a model for the effect of expression of antibiotic binding molecules in Escherichia coli, and for directed evolution towards high antibiotic resistance.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号